Proper reference frame (flat spacetime)

Last updated

A proper reference frame in the theory of relativity is a particular form of accelerated reference frame, that is, a reference frame in which an accelerated observer can be considered as being at rest. It can describe phenomena in curved spacetime, as well as in "flat" Minkowski spacetime in which the spacetime curvature caused by the energy–momentum tensor can be disregarded. Since this article considers only flat spacetime—and uses the definition that special relativity is the theory of flat spacetime while general relativity is a theory of gravitation in terms of curved spacetime—it is consequently concerned with accelerated frames in special relativity. [1] [2] [3] (For the representation of accelerations in inertial frames, see the article Acceleration (special relativity), where concepts such as three-acceleration, four-acceleration, proper acceleration, hyperbolic motion etc. are defined and related to each other.)

Contents

A fundamental property of such a frame is the employment of the proper time of the accelerated observer as the time of the frame itself. This is connected with the clock hypothesis (which is experimentally confirmed), according to which the proper time of an accelerated clock is unaffected by acceleration, thus the measured time dilation of the clock only depends on its momentary relative velocity. The related proper reference frames are constructed using concepts like comoving orthonormal tetrads, which can be formulated in terms of spacetime Frenet–Serret formulas, or alternatively using Fermi–Walker transport as a standard of non-rotation. If the coordinates are related to Fermi–Walker transport, the term Fermi coordinates is sometimes used, or proper coordinates in the general case when rotations are also involved. A special class of accelerated observers follow worldlines whose three curvatures are constant. These motions belong to the class of Born rigid motions, i.e., the motions at which the mutual distance of constituents of an accelerated body or congruence remains unchanged in its proper frame. Two examples are Rindler coordinates or Kottler-Møller coordinates for the proper reference frame of hyperbolic motion, and Born or Langevin coordinates in the case of uniform circular motion.

In the following, Greek indices run over 0,1,2,3, Latin indices over 1,2,3, and bracketed indices are related to tetrad vector fields. The signature of the metric tensor is (-1,1,1,1).

History

Some properties of Kottler-Møller or Rindler coordinates were anticipated by Albert Einstein (1907) [H 1] when he discussed the uniformly accelerated reference frame. While introducing the concept of Born rigidity, Max Born (1909) [H 2] recognized that the formulas for the worldline of hyperbolic motion can be reinterpreted as transformations into a "hyperbolically accelerated reference system". Born himself, as well as Arnold Sommerfeld (1910) [H 3] and Max von Laue (1911) [H 4] used this frame to compute the properties of charged particles and their fields (see Acceleration (special relativity)#History and Rindler coordinates#History). In addition, Gustav Herglotz (1909) [H 5] gave a classification of all Born rigid motions, including uniform rotation and the worldlines of constant curvatures. Friedrich Kottler (1912, 1914) [H 6] introduced the "generalized Lorentz transformation" for proper reference frames or proper coordinates (German : Eigensystem, Eigenkoordinaten) by using comoving Frenet–Serret tetrads, and applied this formalism to Herglotz' worldlines of constant curvatures, particularly to hyperbolic motion and uniform circular motion. Herglotz' formulas were also simplified and extended by Georges Lemaître (1924). [H 7] The worldlines of constant curvatures were rediscovered by several author, for instance, by Vladimír Petrův (1964), [4] as "timelike helices" by John Lighton Synge (1967) [5] or as "stationary worldlines" by Letaw (1981). [6] The concept of proper reference frame was later reintroduced and further developed in connection with Fermi–Walker transport in the textbooks by Christian Møller (1952) [7] or Synge (1960). [8] An overview of proper time transformations and alternatives was given by Romain (1963), [9] who cited the contributions of Kottler. In particular, Misner & Thorne & Wheeler (1973) [10] combined Fermi–Walker transport with rotation, which influenced many subsequent authors. Bahram Mashhoon (1990, 2003) [11] analyzed the hypothesis of locality and accelerated motion. The relations between the spacetime Frenet–Serret formulas and Fermi–Walker transport was discussed by Iyer & C. V. Vishveshwara (1993), [12] Johns (2005) [13] or Bini et al. (2008) [14] and others. A detailed representation of "special relativity in general frames" was given by Gourgoulhon (2013). [15]

Comoving tetrads

Spacetime Frenet–Serret equations

For the investigation of accelerated motions and curved worldlines, some results of differential geometry can be used. For instance, the Frenet–Serret formulas for curves in Euclidean space have already been extended to arbitrary dimensions in the 19th century, and can be adapted to Minkowski spacetime as well. They describe the transport of an orthonormal basis attached to a curved worldline, so in four dimensions this basis can be called a comoving tetrad or vierbein (also called vielbein, moving frame, frame field, local frame, repère mobile in arbitrary dimensions): [16] [17] [18] [19]

 

 

 

 

(1)

Here, is the proper time along the worldline, the timelike field is called the tangent that corresponds to the four-velocity, the three spacelike fields are orthogonal to and are called the principal normal , the binormal and the trinormal . The first curvature corresponds to the magnitude of four-acceleration (i.e., proper acceleration), the other curvatures and are also called torsion and hypertorsion.

Fermi–Walker transport and proper transport

While the Frenet–Serret tetrad can be rotating or not, it is useful to introduce another formalism in which non-rotational and rotational parts are separated. This can be done using the following equation for proper transport [20] or generalized Fermi transport [21] of tetrad , namely [10] [12] [22] [21] [20] [23]

 

 

 

 

(2)

where

or together in simplified form:

with as four-velocity and as four-acceleration, and "" indicates the dot product and "" the wedge product. The first part represents Fermi–Walker transport, [13] which is physically realized when the three spacelike tetrad fields don't change their orientation with respect to the motion of a system of three gyroscopes. Thus Fermi–Walker transport can be seen as a standard of non-rotation. The second part consists of an antisymmetric second rank tensor with as the angular velocity four-vector and as the Levi-Civita symbol. It turns out that this rotation matrix only affects the three spacelike tetrad fields, thus it can be interpreted as the spatial rotation of the spacelike fields of a rotating tetrad (such as a Frenet–Serret tetrad) with respect to the non-rotating spacelike fields of a Fermi–Walker tetrad along the same world line.

Deriving Fermi–Walker tetrads from Frenet–Serret tetrads

Since and on the same worldline are connected by a rotation matrix, it is possible to construct non-rotating Fermi–Walker tetrads using rotating Frenet–Serret tetrads, [24] [25] which not only works in flat spacetime but for arbitrary spacetimes as well, even though the practical realization can be hard to achieve. [26] For instance, the angular velocity vector between the respective spacelike tetrad fields and can be given in terms of torsions and : [12] [13] [27] [28]

and

 

 

 

 

(3a)

Assuming that the curvatures are constant (which is the case in helical motion in flat spacetime, or in the case of stationary axisymmetric spacetimes), one then proceeds by aligning the spacelike Frenet–Serret vectors in the plane by constant counter-clockweise rotation, then the resulting intermediary spatial frame is constantly rotated around the axis by the angle , which finally gives the spatial Fermi–Walker frame (note that the timelike field remains the same): [25]

 

 

 

 

(3b)

For the special case and , it follows and and , therefore ( 3b ) is reduced to a single constant rotation around the -axis: [29] [30] [31] [24]

 

 

 

 

(3c)

Proper coordinates or Fermi coordinates

In flat spacetime, an accelerated object is at any moment at rest in a momentary inertial frame , and the sequence of such momentary frames which it traverses corresponds to a successive application of Lorentz transformations , where is an external inertial frame and the Lorentz transformation matrix. This matrix can be replaced by the proper time dependent tetrads defined above, and if is the time track of the particle indicating its position, the transformation reads: [32]

 

 

 

 

(4a)

Then one has to put by which is replaced by and the timelike field vanishes, therefore only the spacelike fields are present anymore. Subsequently, the time in the accelerated frame is identified with the proper time of the accelerated observer by . The final transformation has the form [33] [34] [35] [36]

,

 

 

 

 

(4b)

These are sometimes called proper coordinates, and the corresponding frame is the proper reference frame. [20] They are also called Fermi coordinates in the case of Fermi–Walker transport [37] (even though some authors use this term also in the rotational case [38] ). The corresponding metric has the form in Minkowski spacetime (without Riemannian terms): [39] [40] [41] [42] [43] [44] [45] [46]

 

 

 

 

(4c)

However, these coordinates are not globally valid, but are restricted to [43]

 

 

 

 

(4d)

Proper reference frames for timelike helices

In case all three Frenet–Serret curvatures are constant, the corresponding worldlines are identical to those that follow from the Killing motions in flat spacetime. They are of particular interest since the corresponding proper frames and congruences satisfy the condition of Born rigidity, that is, the spacetime distance of two neighbouring worldlines is constant. [47] [48] These motions correspond to "timelike helices" or "stationary worldlines", and can be classified into six principal types: two with zero torsions (uniform translation, hyperbolic motion) and four with non-zero torsions (uniform rotation, catenary, semicubical parabola, general case): [49] [50] [4] [5] [6] [51] [52] [53] [54]

Case produces uniform translation without acceleration. The corresponding proper reference frame is therefore given by ordinary Lorentz transformations. The other five types are:

Hyperbolic motion

The curvatures , where is the constant proper acceleration in the direction of motion, produce hyperbolic motion because the worldline in the Minkowski diagram is a hyperbola: [55] [56] [57] [58] [59] [60]

 

 

 

 

(5a)

The corresponding orthonormal tetrad is identical to an inverted Lorentz transformation matrix with hyperbolic functions as Lorentz factor and as proper velocity and as rapidity (since the torsions and are zero, the Frenet–Serret formulas and Fermi–Walker formulas produce the same tetrad): [56] [61] [62] [63] [64] [65] [66]

 

 

 

 

(5b)

Inserted into the transformations ( 4b ) and using the worldline ( 5a ) for , the accelerated observer is always located at the origin, so the Kottler-Møller coordinates follow [67] [68] [62] [69] [70]

which are valid within , with the metric

.

Alternatively, by setting the accelerated observer is located at at time , thus the Rindler coordinates follow from ( 4b ) and ( 5a , 5b ): [71] [72] [73]

which are valid within , with the metric

Uniform circular motion

The curvatures , produce uniform circular motion, with the worldline [74] [75] [76] [77] [78] [79] [80]

 

 

 

 

(6a)

where

 

 

 

 

(6b)

with as orbital radius, as coordinate angular velocity, as proper angular velocity, as tangential velocity, as proper velocity, as Lorentz factor, and as angle of rotation. The tetrad can be derived from the Frenet–Serret equations ( 1 ), [74] [76] [77] [80] or more simply be obtained by a Lorentz transformation of the tetrad of ordinary rotating coordinates: [81] [82]

 

 

 

 

(6c)

The corresponding non-rotating Fermi–Walker tetrad on the same worldline can be obtained by solving the Fermi–Walker part of equation ( 2 ). [83] [84] Alternatively, one can use ( 6b ) together with ( 3a ), which gives

The resulting angle of rotation together with ( 6c ) can now be inserted into ( 3c ), by which the Fermi–Walker tetrad follows [31] [24]

In the following, the Frenet–Serret tetrad is used to formulate the transformation. Inserting ( 6c ) into the transformations ( 4b ) and using the worldline ( 6a ) for gives the coordinates [74] [76] [85] [86] [87] [38]

 

 

 

 

(6d)

which are valid within , with the metric

If an observer resting in the center of the rotating frame is chosen with , the equations reduce to the ordinary rotational transformation [88] [89] [90]

 

 

 

 

(6e)

which are valid within , and the metric

.

The last equations can also be written in rotating cylindrical coordinates (Born coordinates): [91] [92] [93] [94] [95]

 

 

 

 

(6f)

which are valid within , and the metric

Frames ( 6d , 6e , 6f ) can be used to describe the geometry of rotating platforms, including the Ehrenfest paradox and the Sagnac effect.

Catenary

The curvatures , produce a catenary, i.e., hyperbolic motion combined with a spacelike translation [96] [97] [98] [99] [100] [101] [102]

 

 

 

 

(7a)

where

 

 

 

 

(7b)

where is the velocity, the proper velocity, as rapidity, is the Lorentz factor. The corresponding Frenet–Serret tetrad is: [97] [99]

The corresponding non-rotating Fermi–Walker tetrad on the same worldline can be obtained by solving the Fermi–Walker part of equation ( 2 ). [102] The same result follows from ( 3a ), which gives

which together with ( 7a ) can now be inserted into ( 3c ), resulting in the Fermi–Walker tetrad

The proper coordinates or Fermi coordinates follow by inserting or into ( 4b ).

Semicubical parabola

The curvatures , produce a semicubical parabola or cusped motion [103] [104] [105] [106] [107] [108] [109]

with

 

 

 

 

(8)

The corresponding Frenet–Serret tetrad with is: [104] [106]

The corresponding non-rotating Fermi–Walker tetrad on the same worldline can be obtained by solving the Fermi–Walker part of equation ( 2 ). [109] The same result follows from ( 3a ), which gives

which together with ( 8 ) can now be inserted into ( 3c ), resulting in the Fermi–Walker tetrad (note that in this case):

The proper coordinates or Fermi coordinates follow by inserting or into ( 4b ).

General case

The curvatures , , produce hyperbolic motion combined with uniform circular motion. The worldline is given by [110] [111] [112] [113] [114] [115] [116]

 

 

 

 

(9a)

where

 

 

 

 

(9b)

with as tangential velocity, as proper tangential velocity, as rapidity, as orbital radius, as coordinate angular velocity, as proper angular velocity, as angle of rotation, is the Lorentz factor. The Frenet–Serret tetrad is [111] [113]

The corresponding non-rotating Fermi–Walker tetrad on the same worldline is as follows: First inserting ( 9b ) into ( 3a ) gives the angular velocity, which together with ( 9a ) can now be inserted into ( 3b , left), and finally inserted into ( 3b , right) produces the Fermi–Walker tetrad. The proper coordinates or Fermi coordinates follow by inserting or into ( 4b ) (the resulting expressions are not indicated here because of their length).

Overview of historical formulas

In addition to the things described in the previous #History section, the contributions of Herglotz, Kottler, and Møller are described in more detail, since these authors gave extensive classifications of accelerated motion in flat spacetime.

Herglotz

Herglotz (1909) [H 5] argued that the metric

where

satisfies the condition of Born rigidity when . He pointed out that the motion of a Born rigid body is in general determined by the motion of one of its point (class A), with the exception of those worldlines whose three curvatures are constant, thus representing a helix (class B). For the latter, Herglotz gave the following coordinate transformation corresponding to the trajectories of a family of motions:

(H1) ,

where and are functions of proper time . By differentiation with respect to , and assuming as constant, he obtained

(H2)

Here, represents the four-velocity of the origin of , and is a six-vector (i.e., an antisymmetric four-tensor of second order, or bivector, having six independent components) representing the angular velocity of around . As any six-vector, it has two invariants:

When is constant and is variable, any family of motions described by (H1) forms a group and is equivalent to an equidistant family of curves, thus satisfying Born rigidity because they are rigidly connected with . To derive such a group of motion, (H2) can be integrated with arbitrary constant values of and . For rotational motions, this results in four groups depending on whether the invariants or are zero or not. These groups correspond to four one-parameter groups of Lorentz transformations, which were already derived by Herglotz in a previous section on the assumption, that Lorentz transformations (being rotations in ) correspond to hyperbolic motions in . The latter have been studied in the 19th century, and were categorized by Felix Klein into loxodromic, elliptic, hyperbolic, and parabolic motions (see also Möbius group).

Kottler

Friedrich Kottler (1912) [H 6] followed Herglotz, and derived the same worldlines of constant curvatures using the following Frenet–Serret formulas in four dimensions, with as comoving tetrad of the worldline, and as the three curvatures

corresponding to ( 1 ). Kottler pointed out that the tetrad can be seen as a reference frame for such worldlines. Then he gave the transformation for the trajectories

(with )

in agreement with ( 4a ). Kottler also defined a tetrad whose basis vectors are fixed in normal space and therefore do not share any rotation. This case was further differentiated into two cases: If the tangent (i.e., the timelike) tetrad field is constant, then the spacelike tetrads fields can be replaced by who are "rigidly" connected with the tangent, thus

The second case is a vector "fixed" in normal space by setting . Kottler pointed out that this corresponds to class B given by Herglotz (which Kottler calls "Born's body of second kind")

,

and class (A) of Herglotz (which Kottler calls "Born's body of first kind") is given by

which both correspond to formula ( 4b ).


In (1914a), [H 6] Kottler showed that the transformation

,

describes the non-simultaneous coordinates of the points of a body, while the transformation with

,

describes the simultaneous coordinates of the points of a body. These formulas become "generalized Lorentz transformations" by inserting

thus

in agreement with ( 4b ). He introduced the terms "proper coordinates" and "proper frame" (German : Eigenkoordinaten, Eigensystem) for a system whose time axis coincides with the respective tangent of the worldline. He also showed that the Born rigid body of second kind, whose worldlines are defined by

,

is particularly suitable for defining a proper frame. Using this formula, he defined the proper frames for hyperbolic motion (free fall) and for uniform circular motion:

Hyperbolic motionUniform circular motion
1914b1914a1921

In (1916a) Kottler gave the general metric for acceleration-relative motions based on the three curvatures

In (1916b) he gave it the form:

where are free from , and , and , and linear in .

Møller

Møller (1952) [7] defined the following transport equation

in agreement with Fermi–Walker transport by ( 2 , without rotation). The Lorentz transformation into a momentary inertial frame was given by him as

in agreement with ( 4a ). By setting , and , he obtained the transformation into the "relativistic analogue of a rigid reference frame"

in agreement with the Fermi coordinates ( 4b ), and the metric

in agreement with the Fermi metric ( 4c ) without rotation. He obtained the Fermi–Walker tetrads and Fermi frames of hyperbolic motion and uniform circular motion (some formulas for hyperbolic motion were already derived by him in 1943):

Hyperbolic motionUniform circular motion
194319521952

Worldlines of constant curvatures by Herglotz and Kottler

General caseUniform rotationCatenarySemicubical parabolaHyperbolic motion
Herglotz (1909)
loxodromicelliptichyperbolicparabolichyperbolic
Lorentz-Transformations
Trajectories (time)
Kottler (1912, 1914)
hyperspherical curveuniform rotationcatenarycubic curvehyperbolic motion
Curvatures
Trajectory of
Trajectory (time)

Related Research Articles

<span class="mw-page-title-main">Four-vector</span> 4-dimensional vector in relativity

In special relativity, a four-vector is an object with four components, which transform in a specific way under Lorentz transformations. Specifically, a four-vector is an element of a four-dimensional vector space considered as a representation space of the standard representation of the Lorentz group, the representation. It differs from a Euclidean vector in how its magnitude is determined. The transformations that preserve this magnitude are the Lorentz transformations, which include spatial rotations and boosts.

<span class="mw-page-title-main">Theta function</span> Special functions of several complex variables

In mathematics, theta functions are special functions of several complex variables. They show up in many topics, including Abelian varieties, moduli spaces, quadratic forms, and solitons. As Grassmann algebras, they appear in quantum field theory.

<span class="mw-page-title-main">Hamilton–Jacobi equation</span> A reformulation of Newtons laws of motion using the calculus of variations

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

In mathematics and physics, the Christoffel symbols are an array of numbers describing a metric connection. The metric connection is a specialization of the affine connection to surfaces or other manifolds endowed with a metric, allowing distances to be measured on that surface. In differential geometry, an affine connection can be defined without reference to a metric, and many additional concepts follow: parallel transport, covariant derivatives, geodesics, etc. also do not require the concept of a metric. However, when a metric is available, these concepts can be directly tied to the "shape" of the manifold itself; that shape is determined by how the tangent space is attached to the cotangent space by the metric tensor. Abstractly, one would say that the manifold has an associated (orthonormal) frame bundle, with each "frame" being a possible choice of a coordinate frame. An invariant metric implies that the structure group of the frame bundle is the orthogonal group O(p, q). As a result, such a manifold is necessarily a (pseudo-)Riemannian manifold. The Christoffel symbols provide a concrete representation of the connection of (pseudo-)Riemannian geometry in terms of coordinates on the manifold. Additional concepts, such as parallel transport, geodesics, etc. can then be expressed in terms of Christoffel symbols.

<span class="mw-page-title-main">Directional statistics</span>

Directional statistics is the subdiscipline of statistics that deals with directions, axes or rotations in Rn. More generally, directional statistics deals with observations on compact Riemannian manifolds including the Stiefel manifold.

<span class="mw-page-title-main">Hyperbolic motion (relativity)</span> Motion of an object with constant proper acceleration in special relativity.

Hyperbolic motion is the motion of an object with constant proper acceleration in special relativity. It is called hyperbolic motion because the equation describing the path of the object through spacetime is a hyperbola, as can be seen when graphed on a Minkowski diagram whose coordinates represent a suitable inertial (non-accelerated) frame. This motion has several interesting features, among them that it is possible to outrun a photon if given a sufficient head start, as may be concluded from the diagram.

In physics and astronomy, Euler's three-body problem is to solve for the motion of a particle that is acted upon by the gravitational field of two other point masses that are fixed in space. This problem is exactly solvable, and yields an approximate solution for particles moving in the gravitational fields of prolate and oblate spheroids. This problem is named after Leonhard Euler, who discussed it in memoirs published in 1760. Important extensions and analyses were contributed subsequently by Lagrange, Liouville, Laplace, Jacobi, Darboux, Le Verrier, Velde, Hamilton, Poincaré, Birkhoff and E. T. Whittaker, among others.

In probability and statistics, a circular distribution or polar distribution is a probability distribution of a random variable whose values are angles, usually taken to be in the range [0, 2π). A circular distribution is often a continuous probability distribution, and hence has a probability density, but such distributions can also be discrete, in which case they are called circular lattice distributions. Circular distributions can be used even when the variables concerned are not explicitly angles: the main consideration is that there is not usually any real distinction between events occurring at the lower or upper end of the range, and the division of the range could notionally be made at any point.

<span class="mw-page-title-main">Osculating circle</span> Circle of immediate corresponding curvature of a curve at a point

In differential geometry of curves, the osculating circle of a sufficiently smooth plane curve at a given point p on the curve has been traditionally defined as the circle passing through p and a pair of additional points on the curve infinitesimally close to p. Its center lies on the inner normal line, and its curvature defines the curvature of the given curve at that point. This circle, which is the one among all tangent circles at the given point that approaches the curve most tightly, was named circulus osculans by Leibniz.

A theoretical motivation for general relativity, including the motivation for the geodesic equation and the Einstein field equation, can be obtained from special relativity by examining the dynamics of particles in circular orbits about the Earth. A key advantage in examining circular orbits is that it is possible to know the solution of the Einstein Field Equation a priori. This provides a means to inform and verify the formalism.

The Newman–Penrose (NP) formalism is a set of notation developed by Ezra T. Newman and Roger Penrose for general relativity (GR). Their notation is an effort to treat general relativity in terms of spinor notation, which introduces complex forms of the usual variables used in GR. The NP formalism is itself a special case of the tetrad formalism, where the tensors of the theory are projected onto a complete vector basis at each point in spacetime. Usually this vector basis is chosen to reflect some symmetry of the spacetime, leading to simplified expressions for physical observables. In the case of the NP formalism, the vector basis chosen is a null tetrad: a set of four null vectors—two real, and a complex-conjugate pair. The two real members often asymptotically point radially inward and radially outward, and the formalism is well adapted to treatment of the propagation of radiation in curved spacetime. The Weyl scalars, derived from the Weyl tensor, are often used. In particular, it can be shown that one of these scalars— in the appropriate frame—encodes the outgoing gravitational radiation of an asymptotically flat system.

<span class="mw-page-title-main">Proper acceleration</span> Physical acceleration experienced by an object

In relativity theory, proper acceleration is the physical acceleration experienced by an object. It is thus acceleration relative to a free-fall, or inertial, observer who is momentarily at rest relative to the object being measured. Gravitation therefore does not cause proper acceleration, because the same gravity acts equally on the inertial observer. As a consequence, all inertial observers always have a proper acceleration of zero.

In mathematical physics, spacetime algebra (STA) is a name for the Clifford algebra Cl1,3(R), or equivalently the geometric algebra G(M4). According to David Hestenes, spacetime algebra can be particularly closely associated with the geometry of special relativity and relativistic spacetime.

In the differential geometry of surfaces, a Darboux frame is a natural moving frame constructed on a surface. It is the analog of the Frenet–Serret frame as applied to surface geometry. A Darboux frame exists at any non-umbilic point of a surface embedded in Euclidean space. It is named after French mathematician Jean Gaston Darboux.

In classical mechanics, a Liouville dynamical system is an exactly solvable dynamical system in which the kinetic energy T and potential energy V can be expressed in terms of the s generalized coordinates q as follows:

Bilinear time–frequency distributions, or quadratic time–frequency distributions, arise in a sub-field of signal analysis and signal processing called time–frequency signal processing, and, in the statistical analysis of time series data. Such methods are used where one needs to deal with a situation where the frequency composition of a signal may be changing over time; this sub-field used to be called time–frequency signal analysis, and is now more often called time–frequency signal processing due to the progress in using these methods to a wide range of signal-processing problems.

<span class="mw-page-title-main">Gravitational lensing formalism</span>

In general relativity, a point mass deflects a light ray with impact parameter by an angle approximately equal to

In applied mathematics, the Biot–Tolstoy–Medwin (BTM) diffraction model describes edge diffraction. Unlike the uniform theory of diffraction (UTD), BTM does not make the high frequency assumption. BTM sees use in acoustic simulations.

In Einstein's theory of general relativity, the interior Schwarzschild metric is an exact solution for the gravitational field in the interior of a non-rotating spherical body which consists of an incompressible fluid and has zero pressure at the surface. This is a static solution, meaning that it does not change over time. It was discovered by Karl Schwarzschild in 1916, who earlier had found the exterior Schwarzschild metric.

Accelerations in special relativity (SR) follow, as in Newtonian Mechanics, by differentiation of velocity with respect to time. Because of the Lorentz transformation and time dilation, the concepts of time and distance become more complex, which also leads to more complex definitions of "acceleration". SR as the theory of flat Minkowski spacetime remains valid in the presence of accelerations, because general relativity (GR) is only required when there is curvature of spacetime caused by the energy–momentum tensor. However, since the amount of spacetime curvature is not particularly high on Earth or its vicinity, SR remains valid for most practical purposes, such as experiments in particle accelerators.

References

  1. Misner & Thorne & Wheeler (1973), p. 163: "Accelerated motion and accelerated observers can be analyzed using special relativity."
  2. Koks (2006), p. 234. "It is sometimes said that to describe physics properly in an accelerated frame, special relativity is insufficient, and the full machinery of general relativity is necessary for the job. This is quite wrong. Special relativity is entirely sufficient to derive the physics of an accelerated frame."
  3. In some textbooks the same formulas and results for flat spacetime are discussed in the framework of GR, using the historical definition that SR is restricted to inertial frames, while accelerated frames belong to the framework of GR. However, since the results are the same in terms of flat spacetime, it does not affect the content of this article. For instance, Møller (1952) discusses successive Lorentz transformations, successive inertial frames, and tetrad transport (now called Fermi–Walker transport) in §§ 46, 47 related to special relativity, while rigid references frames are discussed in section §§ 90, 96 related to general relativity.
  4. 1 2 Petruv (1964)
  5. 1 2 Synge (1967)
  6. 1 2 Letaw (1981)
  7. 1 2 Møller (1952), §§ 46, 47, 90, 96
  8. Synge (1960), §§ 3, 4
  9. Romain (1963), particularly section VI for the "proper time approach"
  10. 1 2 Misner & Thorne & Wheeler (1973), section 6.8
  11. Mashhoon (1990), (2003)
  12. 1 2 3 Iyer and Vishveshwara (1993), section 2.2
  13. 1 2 3 Johns (2005), section 18.18
  14. Bini & Cherubini & Geralico & Jantzen (2008), section 3
  15. Gourgoulhon (2013)
  16. Synge (1960), § 3
  17. Iyer and Vishveshwara (1993), section 2.1
  18. Formiga & Romero (2006), section 2
  19. Gourgoulhon (2013), section 2.7.3
  20. 1 2 3 Kajari & Buser & Feiler & Schleich (2009), section 3
  21. 1 2 Hehl & Lemke & Mielke (1990), section I.6
  22. Padmanabhan (2010), section 4.9
  23. Gourgoulhon (2013), section 3.5.3
  24. 1 2 3 Johns (2005), section 18.19
  25. 1 2 Bini & Cherubini & Geralico & Jantzen (2008), section 3.2
  26. Maluf & Faria (2008)
  27. Bini & Cherubini & Geralico & Jantzen (2008), section 3.1
  28. Gourgoulhon (2013), eq. 3.58
  29. Irvine (1964), section VII, eq. 41
  30. Bini & Jantzen (2003), Appendix A
  31. 1 2 Mashhoon (2003), section 3, eq. 1.17, 1.18
  32. Møller (1952), § 46
  33. Møller (1952), § 96
  34. Hehl & Lemke & Mielke (1990), section I.8
  35. Mashhoon & Muench (2002), section 2
  36. Kopeikin & Efroimsky & Kaplan (2011), section 2.6
  37. Synge (1960), § 10
  38. 1 2 Bini & Lusanna & Mashhoon (2005), Appendix A
  39. Ni & Zimmermann (1978), including Riemannian terms
  40. Hehl & Lemke & Mielke (1990), section I.8, without Riemannian terms
  41. Marzlin (1994), section 2, including Riemannian terms
  42. Nikolić (1999), section 2, without Riemannian terms
  43. 1 2 Mashhoon & Münch (2002), section 2, without Riemannian terms
  44. Bini & Jantzen (2002), section 2, including Riemannian terms
  45. Voytik (2011), section 2, without Riemannian terms
  46. Misner & Thorne & Wheeler (1973), section 13.6, gave the first order approximation to this metric, without Riemannian terms
  47. Bel (1995), theorem 2
  48. Giulini (2008), Theorem 18
  49. Herglotz (1909), sections 3-4, who focuses on the four rotational motions in addition to hyperbolic motion.
  50. Kottler (1912), § 6; (1914a), table I & II
  51. Letaw & Pfautsch (1982)
  52. Pauri & Vallisneri (2001), Appendix A
  53. Rosu (2000), section 0.2.3
  54. Louko & Satz (2006), section 5.2
  55. Herglotz (1909), p. 408
  56. 1 2 Kottler (1914a), table I (IIIb); Kottler (1914b), pp. 488-489, 492-493
  57. Petruv (1964), eq. 22
  58. Synge (1967), section 9
  59. Pauri & Vallisneri (2001), eq. 19
  60. Rosu (2000), section 0.2.3, case 2
  61. Møller (1952) eq. 160
  62. 1 2 Synge (1967) p. 35, type III
  63. Misner & Thorne & Wheeler (1973), section 6.4
  64. Louko & Satz (2006), section 5.2.2
  65. Gron (2006), section 5.5
  66. Formiga (2012), section V-a
  67. Kottler (1914b), pp. 488-489, 492-493
  68. Møller (1952), eq. 154
  69. Misner & Thorne & Wheeler (1973), section 6.6
  70. Muñoz & Jones (2010), eq. 37, 38
  71. Pauli (1921), section 32-y
  72. Rindler (1966), p. 1177
  73. Koks (2006), section 7.2
  74. 1 2 3 Kottler (1914a), table I (IIb) and § 6 section 3
  75. Petruv (1964), eq. 54
  76. 1 2 3 Nožička (1964), example 1
  77. 1 2 Synge (1967), section 8
  78. Pauri & Vallisneri (2001), eq. 20
  79. Rosu (2000), section 0.2.3, case 3
  80. 1 2 Formiga (2012), section V-b
  81. Hauck & Mashhoon (2003), section 1
  82. Mashhoon (2003), section 3
  83. Møller (1952), § 47, eq. 164
  84. Louko & Satz (2006), section 5.2.3
  85. Mashhoon (1990), eq. 10-13
  86. Nikolic (1999), eq. 17 (He obtained these formulas by using the transformation of Nelson).
  87. Mashhoon (2003), eq. 1.22-1.25
  88. Herglotz (1909), p. 412, "elliptic group"
  89. Eddington (1920), p. 22.
  90. de Felice (2003), section 2
  91. de Sitter (1916a), p. 178
  92. von Laue (1921), p. 162
  93. Gron (2006), section 5.1
  94. Rizzi & Ruggiero (2002), p. section 5
  95. Ashby (2003), section 2
  96. Herglotz (1909), pp. 408 & 413, "hyperbolic group"
  97. 1 2 Kottler (1914a), table I (IIIa)
  98. Petruv (1964), eq. 67
  99. 1 2 Synge (1967), section 6
  100. Pauri & Vallisneri (2001), eq. 22
  101. Rosu (2000), section 0.2.3, case 5
  102. 1 2 Louko & Satz (2006), section 5.2.5
  103. Herglotz (1909), pp. 413-414, "parabolic group"
  104. 1 2 Kottler (1914a), table I (IV)
  105. Petruv (1964), eq. 40
  106. 1 2 Synge (1967), section 7
  107. Pauri & Vallisneri (2001), eq. 21
  108. Rosu (2000), section 0.2.3, case 4
  109. 1 2 Louko & Satz (2006), section 5.2.4
  110. Herglotz (1909), pp. 411-412, "parabolic group"
  111. 1 2 Kottler (1914a), table I (case I)
  112. Petruv (1964), eq. 88
  113. 1 2 Synge (1967), section 4
  114. Pauri & Vallisneri (2001), eq. 23, 24
  115. Rosu (2000), section 0.2.3, case 6
  116. Louko & Satz (2006), section 5.2.6

Bibliography

Textbooks

Journal articles

Historical sources

  1. Einstein, Albert (1908) [1907], "Über das Relativitätsprinzip und die aus demselben gezogenen Folgerungen" (PDF), Jahrbuch der Radioaktivität und Elektronik, 4: 411–462, Bibcode:1908JRE.....4..411E ; English translation On the relativity principle and the conclusions drawn from it at Einstein paper project.
  2. Born, Max (1909), "Die Theorie des starren Elektrons in der Kinematik des Relativitätsprinzips" [Wikisource translation: The Theory of the Rigid Electron in the Kinematics of the Principle of Relativity ], Annalen der Physik, 335 (11): 1–56, Bibcode:1909AnP...335....1B, doi:10.1002/andp.19093351102
  3. Sommerfeld, Arnold (1910). "Zur Relativitätstheorie II: Vierdimensionale Vektoranalysis" [Wikisource translation: On the Theory of Relativity II: Four-dimensional Vector Analysis ]. Annalen der Physik. 338 (14): 649–689. Bibcode:1910AnP...338..649S. doi:10.1002/andp.19103381402.
  4. Laue, Max von (1911). Das Relativitätsprinzip. Braunschweig: Vieweg.
  5. 1 2 Herglotz, Gustav (1910) [1909], "Über den vom Standpunkt des Relativitätsprinzips aus als starr zu bezeichnenden Körper" [Wikisource translation: On bodies that are to be designated as "rigid" from the standpoint of the relativity principle ], Annalen der Physik, 336 (2): 393–415, Bibcode:1910AnP...336..393H, doi:10.1002/andp.19103360208
  6. 1 2 3 Kottler, Friedrich (1912). "Über die Raumzeitlinien der Minkowski'schen Welt" [Wikisource translation: On the spacetime lines of a Minkowski world ]. Wiener Sitzungsberichte 2a. 121: 1659–1759. hdl:2027/mdp.39015051107277.Kottler, Friedrich (1914a). "Relativitätsprinzip und beschleunigte Bewegung". Annalen der Physik. 349 (13): 701–748. Bibcode:1914AnP...349..701K. doi:10.1002/andp.19143491303.Kottler, Friedrich (1914b). "Fallende Bezugssysteme vom Standpunkte des Relativitätsprinzips". Annalen der Physik. 350 (20): 481–516. Bibcode:1914AnP...350..481K. doi:10.1002/andp.19143502003.Kottler, Friedrich (1916a). "Beschleunigungsrelative Bewegungen und die konforme Gruppe der Minkowski'schen Welt". Wiener Sitzungsberichte 2a. 125: 899–919. hdl:2027/mdp.39015073682984.Kottler, Friedrich (1916b). "Über Einsteins Äquivalenzhypothese und die Gravitation". Annalen der Physik. 355 (16): 955–972. Bibcode:1916AnP...355..955K. doi:10.1002/andp.19163551605.Kottler, Friedrich (1918). "Über die physikalischen Grundlagen der Einsteinschen Relativitätstheorie". Annalen der Physik. 361 (14): 401–461. Bibcode:1918AnP...361..401K. doi:10.1002/andp.19183611402.Kottler, Friedrich (1921). "Rotierende Bezugssysteme in einer Minkowskischen Welt". Physikalische Zeitschrift. 22: 274–280 & 480–484. hdl:2027/mdp.39015020056829.
  7. Lemaître, G. (1924), "The motion of a rigid solid according to the relativity principle", Philosophical Magazine, Series 6, 48 (283): 164–176, doi:10.1080/14786442408634478